arxiv:gr-qc/ v2 23 Aug 1996

Size: px
Start display at page:

Download "arxiv:gr-qc/ v2 23 Aug 1996"

Transcription

1 Quantum Theory of Geometry I: Area Operators arxiv:gr-qc/ v2 23 Aug 1996 Abhay Ashtekar 1 and Jerzy Lewandowski 2,3 1 Center for Gravitational Physics and Geometry Physics Department, Penn State, University Park, PA 16802, USA 2 Institute of Theoretical Physics, Warsaw University, ul Hoza 69, Warsaw, Poland 3 Max Planck Institut für Gravitationphysik, Schlaatzweg 1, Potsdam, Germany Abstract A new functional calculus, developed recently for a fully nonperturbative treatment of quantum gravity, is used to begin a systematic construction of a quantum theory of geometry. Regulated operators corresponding to areas of 2-surfaces are introduced and shown to be self-adjoint on the underlying (kinematical) Hilbert space of states. It is shown that their spectra are purely discrete indicating that the underlying quantum geometry is far from what the continuum picture might suggest. Indeed, the fundamental excitations of quantum geometry are 1-dimensional, rather like polymers, and the 3-dimensional continuum geometry emerges only on coarse graining. The full Hilbert space admits an orthonormal decomposition into finite dimensional sub-spaces which can be interpreted as the spaces of states of spin systems. Using this property, the complete spectrum of the area operators is evaluated. The general framework constructed here will be used in a subsequent paper to discuss 3-dimensional geometric operators, e.g., the ones corresponding to volumes of regions. It is a pleasure to dedicate this article to Professor Andrzej Trautman, who was one of the first to recognize the deep relation between geometry and the physics of gauge fields 1,2 which lies at the heart of this investigation. Electronic address: ashtekar@phys.psu.edu Electronic address: jerzy.lewandowski@fuw.edu.pl 1

2 I. INTRODUCTION In his celebrated inaugural address, Riemann suggested [3] that geometry of space may be more than just a fiducial, mathematical entity serving as a passive stage for physical phenomena, and may in fact have a direct physical meaning in its own right. General relativity proved this vision to be correct: Einstein s equations put geometry on the same footing as matter. Now, the physics of this century has shown us that matter has constituents and the 3-dimensional objects we perceive as solids in fact have a discrete underlying structure. The continuum description of matter is an approximation which succeeds brilliantly in the macroscopic regime but fails hopelessly at the atomic scale. It is therefore natural to ask if the same is true of geometry. Does geometry also have constituents at the Planck scale? What are its atoms? Its elementary excitations? Is the space-time continuum only a coarse-grained approximation? If so, what is the nature of the underlying quantum geometry? To probe such issues, it is natural to look for hints in the procedures that have been successful in describing matter. Let us begin by asking what we mean by quantization of physical quantities. Let us take a simple example the hydrogen atom. In this case, the answer is clear: while the basic observables energy and angular momentum take on a continuous range of values classically, in quantum mechanics their spectra are discrete. So, we can ask if the same is true of geometry. Classical geometrical observables such as areas of surfaces and volumes of regions can take on continuous values on the phase space of general relativity. Are the spectra of corresponding quantum operators discrete? If so, we would say that geometry is quantized. Thus, it is rather easy to pose the basic questions in a precise fashion. Indeed, they could have been formulated soon after the advent of quantum mechanics. Answering them, on the other hand, has proved to be surprisingly difficult. The main reason, it seems, is the inadequacy of the standard techniques. More precisely, the traditional approach to quantum field theory has been perturbative, where one begins with a continuum, background geometry. It is then difficult to see how discreteness would arise in the spectra of geometric operators. To analyze such issues, one needs a fully non-perturbative approach: geometric operators have to be constructed ab-initio without assuming any background geometry. To probe the nature of quantum geometry, we can not begin by assuming the validity of the continuum picture. We must let quantum gravity itself decide whether this picture is adequate at the Planck scale; the theory itself should lead us to the correct microscopic picture of geometry. In this paper, we will use the non-perturbative, canonical approach to quantum gravity based on connections to probe these issues. Over the past three years, this approach has been put on a firm mathematical footing through the development of a new functional calculus on the space of gauge equivalent connections [4-11]. This calculus does not use any background fields (such as a metric) and is therefore well-suited to a fully nonperturbative treatment. The purpose of this paper is to use this framework to explore the nature of quantum geometry. In section 2, we recall the relevant results from the new functional calculus and out- 2

3 line the general strategy. In section 3, we present a regularization of the area operator. Its properties are discussed in section 4; in particular, we exhibit its entire spectrum. Our analysis is carried out in the connection representation and the discussion is selfcontained. However, at a non-technical level, there is a close similarity between the basic ideas used here and those used in discussions based on the loop representation [12,13]. Indeed, the development of the functional calculus which underlies this analysis itself was motivated, in a large measure, by the pioneering work on loop representation by Rovelli and Smolin [14]. The relation between various approaches will discussed in section 5. The main result of this paper should have ramifications on the statistical mechanical origin of the entropy of black holes along the lines of [15,16]. This issue is being investigated. II. PRELIMINARIES This section is divided into three parts. In the first, we will recall [4,5] the basic structure of the quantum configuration space and, in the second, that of the Hilbert space of (kinematic) quantum states [10]. The overall strategy will be summarized in the third part. A. Quantum configuration space In general relativity, one can regard the space A/G of SU(2) connections modulo gauge transformations on a ( spatial ) 3-manifold Σ as the classical configuration space [17 19]. For systems with only a finite number of degrees of freedom, the classical configuration space also serves as the domain space of quantum wave functions, i.e., as the quantum configuration space. For systems with an infinite number of degrees of freedom, on the other hand, this is not true: generically, the quantum configuration space is an enlargement of the classical. In free field theory in Minkowski space (as well as exactly solvable models in low space-time dimensions), for example, while the classical configuration space can be built from suitably smooth fields, the quantum configuration space includes all (tempered) distributions. This is an important point because, typically, the classical configuration spaces are of zero measure; wave functions with support only on smooth configurations have zero norm! The overall situation is the same in general relativity. The quantum configuration space A/G is a certain completion of A/G [4,5]. The space A/G inherits the quotient structure of A/G, i.e., A/G is the quotient of the space A of generalized connections by the space G of generalized gauge transformations. To see the nature of the generalization involved, recall first that each smooth connection defines a holonomy along paths 1 in Σ: h p (A) := P exp p A. Generalized connections 1 For technical reasons, we will assume that all paths are analytic. An extension of the frame- 3

4 capture this notion. That is, each Ā in A can be defined [6,8] as a map which assigns to each oriented path p in Σ an element Ā(p) of SU(2) such that: i) Ā(p 1 ) = (Ā(p)) 1 ; and, ii) Ā(p 2 p 1 ) = Ā(p 2) Ā(p 1), where p 1 is obtained from p by simply reversing the orientation, p 2 p 1 denotes the composition of the two paths (obtained by connecting the end of p 1 with the beginning of p 2 ) and Ā(p 2) Ā(p 1) is the composition in SU(2). A generalized gauge transformation is a map g which assigns to each point v of Σ an SU(2) element g(x) (in an arbitrary, possibly discontinuous fashion). It acts on Ā in the expected manner, at the end points of paths: Ā(p) g(v + ) 1 Ā(p) g(v ), where v and v + are respectively the beginning and the end point of p. If Ā happens to be a smooth connections, say A, we have Ā(p) = h p(a). However, in general, Ā(p) can not be expressed as a path ordered exponential of a smooth 1-form with values in the Lie algebra of SU(2) [5]. Similarly, in general, a generalized gauge transformation can not be represented by a smooth group valued function on Σ. At first sight the spaces A, G and A/G seem too large to be mathematically controllable. However, they admit three characterizations which enables one to introduce differential and integral calculus on them [4,5,7]. We will conclude this sub-section by summarizing the characterization as suitable limits of the corresponding spaces in lattice gauge theory which will be most useful for the main body of this paper. We begin with some definitions. An edge is an oriented, 1-dimensional sub-manifold of Σ with two boundary points, called vertices, which is analytic everywhere, including the vertices. A graph in Σ is a collection of edges such that if two distinct edges meet, they do so only at vertices. In the physics terminology, one can think of a graph as a floating lattice, i.e., a lattice whose edges are not required to be rectangular. (Indeed, they may even be non-trivially knotted!) Using the standard ideas from lattice gauge theory, we can construct the configuration space associated with the graph γ. Thus, we have the space A γ, each element A γ of which assigns to every edge in γ an element of SU(2) and the space G γ each element g γ of which assigns to each vertex in γ an element of SU(2). (Thus, if N is the number of edges in γ and V the number of vertices, A γ is isomorphic with [SU(2)] N and G γ with [SU(2)] V ). G γ has the obvious action on A γ : A γ (e) g(v + ) 1 A γ (e) g(v ). The (gauge invariant) configuration space associated with the floating lattice γ is just A γ /G γ. The spaces A, G and A/G can be obtained as well-defined (projective) limits of the spaces A γ, G γ and A γ /G γ [7,5]. Note however that this limit is not the usual continuum limit of a lattice gauge theory in which one lets the edge length go to zero. Here, we are already in the continuum and have available to us all possible floating lattices from the beginning. We are just expressing the quantum configuration space of the continuum theory as a suitable limit of the configuration spaces of theories associated with all these lattices. work to allow for smooth paths is being carried out [20]. The general expectation is that the main results will admit natural generalizations to the smooth category. In this article, A has the physical dimensions of a connection, (length) 1 and is thus related to the configuration variable A old in the literature by A = GA old where G is Newton s constant. 4

5 To summarize, the quantum configuration space A/G is a specific extension of the classical configuration space A/G. Quantum states can be expressed as complex-valued, square-integrable functions on A/G, or, equivalently, as G-invariant square-integrable functions on A. As in Minkowskian field theories, while A/G is dense in A/G topologically, measure theoretically it is generally sparse; typically, A/G is contained in a subset set of zero measure of A/G [7]. Consequently, what matters is the value of wave functions on genuinely generalized connections. In contrast with the usual Minkowskian situation, however, A, G and A/G are all compact spaces in their natural (Gel fand) topologies [4-8]. This fact simplifies a number of technical issues. Our construction can be compared with the general framework of second quantization proposed by Kijowski [21] already twenty years ago. He introduced the space of states for a field theory by using the projective limit of spaces of states associated to a family of finite dimensional theories. He also, suggested, as an example, the lattice approach. The common element with the present approach is that in our case the space of measures on A is also the projective limit of the spaces of measures defined on finite dimensional spaces A γ. B. Hilbert space Since A/G is compact, it admits regular (Borel, normalized) measures and for every such measure we can construct a Hilbert space of square-integrable functions. Thus, to construct the Hilbert space of quantum states, we need to select a specific measure on A/G. It turns out that A admits a measure µ o that is preferred by both mathematical and physical considerations [5,6]. Mathematically, the measure µ o is natural because its definition does not involve introduction of any additional structure: it is induced on A by the Haar measure on SU(2). More precisely, since A γ is isomorphic to [SU(2)] N, the Haar measure on SU(2) induces on it a measure µ o γ in the obvious fashion. As we vary γ, we obtain a family of measures which turn out to be compatible in an appropriate sense and therefore induce a measure µ o on A. This measure has the following attractive properties [5]: i) it is faithful; i.e., for any continuous, non-negative function f on A, dµ o f 0, equality holding if and only if f is identically zero; and, ii) it is invariant under the (induced) action of Diff[Σ], the diffeomorphism group of Σ. Finally, µ o induces a natural measure µ o on A/G: µ o is simply the push-forward of µ o under the projection map that sends A to A/G. Physically, the measure µ o is selected by the so-called reality conditions. More precisely, the classical phase space admits an (over)complete set of naturally defined configuration and momentum variables which are real, and the requirement that the corresponding operators on the quantum Hilbert space be self-adjoint selects for us the measure µ o [10]. Thus, it is natural to use H o := L 2 (A/G, d µ o ) as our Hilbert space. Elements of Ho are the kinematic states; we are yet to impose quantum constraints. Thus, Ho is the classical analog of the full phase-space of quantum gravity (prior to the introduction of 5

6 the constraint sub-manifold). Note that these quantum states can be regarded also as gauge invariant functions on A. In fact, since the spaces under consideration are compact and measures normalized, we can regard H o as the gauge invariant sub-space of the Hilbert space H o := L 2 (A, dµ o ) of square-integrable functions on A [6,7]. In what follows, we we will often do so. What do typical quantum states look like? To provide an intuitive picture, we can proceed as follows. Fix a graph γ with N edges and consider functions Ψ γ of generalized connections of the form Ψ γ (Ā) = ψ(ā(e 1),..., Ā(e N)) for some smooth function ψ on [SU(2)] N, where e 1,..., e N are the edges of the graph γ. Thus, the functions Ψ γ know about what the generalized connections do only to those paths which constitute the edges of the graph γ; they are precisely the quantum states of the gauge theory associated with the floating lattice γ. This space of states, although infinite dimensional, is quite small in the sense that it corresponds to the Hilbert space associated with a system with only a finite number of degrees of freedom. However, if we vary γ through all possible graphs, the collection of all states that results is very large. Indeed, one can show that it is dense in the Hilbert space H o. (If we restrict ourselves to Ψ γ which are gauge invariant, we obtain a dense sub-space in H o.) Since each of these states depends only on a finite number of variables, borrowing the terminology from the quantum theory of free fields in Minkowski space, they are called cylindrical functions and denoted by Cyl. Gauge invariant cylindrical functions represent the typical kinematic states. In many ways, Cyl is analogous to the space Co (R 3 ) of smooth functions of compact support on R 3 which is dense in the Hilbert space L 2 (R 3, d 3 x) of quantum mechanics. Just as one often defines quantum operators e.g., the position, the momentum and the Hamiltonians on Co first and then extends them to an appropriately larger domain in the Hilbert space L 2 (R 3, d 3 x), we will define our operators first on Cyl and then extend them appropriately. Cylindrical functions provide considerable intuition about the nature of quantum states we are led to consider. These states represent 1-dimensional polymer-like excitations of geometry/gravity rather than 3-dimensional wavy undulations on flat space. Just as a polymer, although intrinsically 1-dimensional, exhibits 3-dimensional properties in sufficiently complex and densely packed configurations, the fundamental 1-dimensional excitations of geometry can be packed appropriately to provide a geometry which, when coarse-grained on scales much larger than the Planck length, lead us to continuum geometries [12,22]. Thus, in this description, gravitons can arise only as approximate notions in the low energy regime [23]. At the basic level, states in H o are fundamentally different from the Fock states of Minkowskian quantum field theories. The main reason is the underlying diffeomorphism invariance: In absence of a background geometry, it is not possible to introduce the familiar Gaussian measures and associated Fock spaces. C. Statement of the problem We can now outline the general strategy that will be followed in sections 4 and 5. 6

7 Recall that the classical configuration variable is an SU(2) connection 2 A i a on a 3- manifold Σ, where i is the su(2)-internal index with respect to a basis τ i. Its conjugate momentum Ej b has the geometrical interpretation of an orthonormal triad with density weight one [24,17], the precise Poisson brackets being: {A i a (x), Eb j (y)} = Gδb a δi j δ3 (x, y), (2.1) where G is Newton s constant. (Recall from footnote 1 that the field A, used here, is related to A old used in the literature [25] via A = GA old.) Therefore, geometrical observables functionals of the 3-metric can be expressed in terms of this field Ei a. Fix within the 3-manifold Σ any analytic, finite 2-surface S without boundary such that the closure of S in Σ is a compact. The area A S of S is a well-defined, real-valued function on the full phase space of general relativity (which happens to depend only on Ei a ). It is easy to verify that these kinematical observables can be expressed as: A S := S dx 1 dx 2 [E 3 i E3i ] 1 2, (2.2) where, for simplicity, we have used adapted coordinates such that S is given by x 3 = 0, and x 1, x 2 parameterize S, and where the internal index i is raised by a the inner product we use on su(2), k(τ i, τ j ) = 2Tr(τ i τ j ). Our task is to find the corresponding operators on the kinematical Hilbert space H o and investigate their properties. There are several factors that make this task difficult. Intuitively, one would expect that Ei a(x) to be replaced by the operator-valued distribution i hgδ/δai a (x). Unfortunately, the classical expression of A S involves square-roots of products of E s and hence the formal expression of the corresponding operator is badly divergent. One must introduce a suitable regularization scheme. Unfortunately, we do not have at our disposal the usual machinery of Minkowskian field theories and even the precise rules that are to underlie such a regularization are not apriori clear. There are however certain basic expectations that we can use as guidelines: i) the resulting operators should be well-defined on a dense sub-space of Ho ; ii) their final expressions should be diffeomorphism covariant, and hence, in particular, independent of any background fields that may be used in the intermediate steps of the regularization procedure; and, iii) since the classical observables are real-valued, the operators should be self-adjoint. These expectations seem to be formidable at first. Indeed, these demands are rarely met even in Minkowskian field theories; in presence of interactions, it is extremely difficult to establish rigorously that physically interesting operators are well-defined and 2 We assume that the underlying 3-manifold Σ is orientable. Hence, principal SU(2) bundles over Σ are all topologically trivial. Therefore, we can represent the SU(2) connections on the bundle by a su(2)-valued 1-form on Σ. The matrices τ i are anti-hermitian, given, e.g., by ( i/2)-times the Pauli matrices. 7

8 self-adjoint. As we will see, the reason why one can succeed in the present case is twofolds. First, the requirement of diffeomorphism covariance is a powerful restriction that severely limits the possibilities. Second, the background independent functional calculus is extremely well-suited for the problem and enables one to circumvent the various road blocks in subtle ways. Our general strategy will be following. We will define the regulated versions of area operators on the dense sub-space Cyl of cylindrical functions and show that they are essentially self-adjoint (i.e., admit unique self-adjoint extensions to H o ). This task is further simplified because the operators leave each sub-space H γ spanned by cylindrical functions associated with any one graph γ invariant. This in effect reduces the field theory problem (i.e., one with an infinite number of degrees of freedom) to a quantum mechanics problem (in which there are only a finite number of degrees of freedom). Finally, we will find that the operators in fact leave invariant certain finite dimensional sub-space of H o (associated with extended spin networks, introduced in Sec. 4.2). This powerful simplification further reduces the task of investigating the properties of these operators; in effect, the quantum mechanical problem (in which the Hilbert space is still infinite dimensional) is further simplified to a problem involving spin systems (where the Hilbert space is finite dimensional). It is because of these simplifications that a complete analysis is possible. III. REGULARIZATION Our task is to construct a well-defined operator ÂS starting from the classical expression (2.2). As is usual in quantum field theory, we will begin with the formal expression obtained by replacing E 3 i in (2.2) by the corresponding operator valued distribution Ê3 i and then regulate it to obtain the required ÂS. (For an early discussion of non-perturbative regularization, see, in particular, [26]). Our discussion will be divided in to two parts. In the first, we introduce the basic tools and, in the second, we apply them to obtain a well-defined operator ÂS. To simplify the presentation, let us first assume that S is covered by a single chart of adapted coordinates. Extension to the general case is straightforward: one mimics the procedure used to define the integral of a differential form over a manifold. That is, one takes advantage of the coordinates invariance of the the resulting local operator and uses a partition of unity. A. Tools The regularization procedure involves two main ingredients. We will begin by summarizing them. The first involves smearing of (the operator analog of) E 3 i (x) and point splitting of the integrand in (2.2). Since in this integrand, the point x lies on the 2-surface S, let 8

9 us try to use a 2-dimensional smearing function. Let f ǫ (x, y) be a 1-parameter family of fields on S which tend to the δ(x, y) as ǫ tends to zero; i.e., such that lim d 2 y f ǫ (x 1, x 2 ; y 1, y 2 )g(y 1, y 2 ) = g(x 1, x 2 ), (3.1) ǫ 0 S for all smooth densities g of weight 1 and of compact support on S. (Thus, f ǫ (x, y) is a density of weight 1 in x and a function in y.) The smeared version of Ei 3 (x) will be defined to be: [Ei 3 ] f (x) := d 2 y f ǫ (x, y)ei 3 (y), (3.2) S so that, as ǫ tends to zero, [Ei 3] f tends to Ei 3 (x). The point-splitting strategy now provides a regularized expression of area: [A S ] f := d 2 x [ d 2 y f ǫ (x, y)ei 3 (y) d 2 z f ǫ (x, z)e 3i (z) ] 1 2 S S S = d 2 x [[Ei 3 ] f(x)[e 3i ] f (x)] 1 2, (3.3) S which will serve as the point of departure in the next subsection. To simplify technicalities, we will assume that the smearing field f ǫ (x, y) has the following additional properties for sufficiently small ǫ > 0: i) for any given y, f ǫ (x, y) has compact support in x which shrinks uniformly to y; and, ii) f ǫ (x, y) is non-negative. These conditions are very mild and we are thus left with a large class of regulators. 3 We now introduce the second ingredient. To go over to the quantum theory, we want to replace Ei 3 in (3.3) by Ê3 i = ig hδ/δa i 3. However, it is not apriori clear that, even after smearing, [Ê3 i ] f is a well-defined operator because: i)our wave functions Ψ are functionals of generalized connections Ā, whence it is not obvious what the functional derivative means; and, ii) we have smeared the operator only along two dimensions. Let us discuss these points one by one. First, let us fix a graph γ and consider a cylindrical function Ψ γ on A, Ψ γ (Ā) = ψ(ā(e 1),.., Ā(e N)), (3.4) where, as before, N is the total number of edges of γ and where ψ is a smooth function on [SU(2)] N. Now, a key fact about generalized connections is that, for any given graph γ, each Ā is equivalent to some smooth connection A [5]: Given any Ā, there exists an A such that Ā(e k ) = h k [A] := P exp A, (3.5) e k 3 For example, f ǫ (x,y) can be constructed as follows. Take any non-negative function f of compact support on S such that d 2 xf(x) = 1 and set f ǫ (x,y) = (1/ǫ 2 )f((x y)/ǫ). Here, we have implicitly used the given chart to give f ǫ (x,y) a density weight in x. 9

10 for all k = 1,..., N. (For any given Ā, the smooth connection A is of course not unique. However, this ambiguity does not affect the considerations that follow.) Hence, there is a 1-1 correspondence between the cylindrical function Ψ γ on A and function ψ(h 1 (A),...,- h E (A)) on the space A of smooth connections and we can apply the operator [Ê3 i ] f to the latter. The result is: N [Ê3 i ] f(x) Ψ γ (Ā) = ig h = il 2 P S I=1 S d 2 y f ǫ (x, y) d 2 δh I yf ǫ (x, y) ( δa i a(y) ) y 3 =0 ( ψ )(A) h I N I=1 (h I (1, t)τ i h I (t, 0)) A B ] ψ 1 [ dt e 3 I (t) δ(y 1, e 1 I (t))δ(y2, e 2 I (t))δ(0, e3 I (t)) 0 h I A B (A), (3.6) where, l P = G h is the Planck length, the index I labels the edges in the graph, [0, 1] t e I (t) is any parameterization of an edge e I, h I (t, t) := P exp t t A a (e I (s))ė a I (s)ds is the holonomy of the connection A along the edge e I from parameter value t to t. Thus, the functional derivative has a well-defined action on cylindrical functions; the first of the two problems mentioned above has been overcome. However, because of the presence of the delta distributions, it is still not clear that [Ê3 i ] f is a genuine operator (rather than a distribution-valued operator). To explicitly see that it is, we need to specify some further details. Given a graph γ, we can just subdivide some of its its edges and thus obtain a graph γ which occupies the same points in Σ as γ but has (trivially) more vertices and edges. Every function which is cylindrical with respect to the smaller graph γ is obviously cylindrical with respect to the larger graph γ as well. The idea is to use this freedom to simplify the discussion by imposing some conditions on our graph γ. We will assume that: i) if an edge e I contains a segment which lies in S, then it lies entirely in the closure of S; ii) each isolated intersection of γ with the 2-surface S is a vertex of γ; and, iii) each edge e I of γ intersects S at most once. (The overlapping edges are often called edges tangential to S; they should not be confused with edges which cross S but whose tangent vector at the intersection point is tangent to S). If the given graph does not satisfy one or more of these conditions, we can obtain one which does simply by sub-dividing of some of the edges. Thus these conditions are not restrictive. They are introduced to simplify the book-keeping in calculations. Let us now return to (3.6). If an edge e I has no point in common with S, it does not contribute to the sum. If it is contained in S, ė 3 I vanishes identically whence its contribution also vanishes. (For a subtlety, see the remark below Eq (3.11).) We are thus left with edges which intersect S in isolated points. Let us first consider only those edges which are outgoing at the intersection. Then, at the intersection point, the value of the parameter t is zero and, for a given edge e I, e 3 I is positive (negative) if e I is directed upwards along increasing x 3 ( downwards along decreasing x 3 ). Hence, (3.6) 10

11 becomes 4 : [Ê3 i ] f(x) Ψ γ = il2 P 2 = il2 P 2 [ N I=1 N I=1 [ d 2 yκ I f ǫ (x, y)δ(y 1, e 1 I (0))δ(y2, e 2 I (0))(h Iτ i ) A B ] ψ A S h IB κ I f ǫ (x, e I (0)) L i I ψ(ā(e 1),..., Ā(e N)), (3.7) where, the constant κ I associated with the edge e I is given by: 0, if e I is tangential to S or does not intersect S, κ I = +1, if e I has an isolated intersection with S and lies above S 1, if e I has an isolated intersection with S and lies below S (3.8) and where L i I is the left invariant vector field in the i-th internal direction on the copy of SU(2) corresponding to the I-th edge L i I ψ(ā(e 1),..., Ā(e N)) = (Ā(e I)τ i ) A B ψ (Ā(e. (3.9) I)) A B If some of the edges are incoming at the intersection point, then the final expression of [Êa i ] f(x) can be written as: [Ê3 i ] f(x) Ψ γ = il2 P 2 [ N κ I f ǫ (x, v αi ) XI] i ψ( Ā(e 1 ),..., Ā(e N)), (3.10) I=1 where XI i is an operator assigned to a vertex v and an edge e I intersecting v by the following formula XI i ψ(ā(e 1),..., Ā(e N)) = (Ā(e I)τ i ) A ψ B, when e (Ā(e I)) A I is outgoing B (τ i Ā(e I )) A (3.11) B, when e I is incoming ψ (Ā(e I)) A B Remark: Let us briefly return to the edges which are tangential to S. In this case, although ė 3 I vanishes, we also have a singular term δ(0, 0) (in the x3 direction) in (3.6). Hence, to recover an unambiguous answer, for these edges, we need to smear also in the third direction using an additional regulator, say g ǫ (x 3, y 3 ). When this is done, one finds that the contribution of the tangential edges vanishes even before removing the regulator; as stated earlier, the tangential edges do not contribute. We did not introduce the smearing in the third direction right in the beginning to emphasize the point that this step is unnecessary for the edges whose contributions survive in the end. The right side again defines a cylindrical function based on the (same) graph γ. Denote by Hγ o the Hilbert space L2 (A γ, dµ o γ ) of square integrable cylindrical functions associated 4 In the first step, we have used the regularization 0 dzg(z)δ(z) = 1 2g(0) which follows if the δ(z) is obtained, in the standard fashion, as a limit of functions which are symmetric about 0. 11

12 with a fixed graph γ. Since µ o γ is the induced Haar measure on A γ and since the operator is just a sum of right/left invariant vector fields, standard results in analysis imply that, with domain Cyl 1 γ of all C 1 cylindrical functions based on γ, it is an essentially self-adjoint on Hγ o. Now, it is straightforward to verify that the operators on Ho γ obtained by varying γ are all compatible 5 in the appropriate sense. Hence, it follows from the general results in [8] that [Ê3 i ] f(x), with domain Cyl 1 (the space of all C 1 cylindrical functions), is an essentially self-adjoint operator on H o. For notational simplicity, we will denote its selfadjoint extension also by [Ê3 i ] f (x). (The context should make it clear whether we are referring to the essentially self-adjoint operator or its extension.) The fact that this operator is well-defined may seem surprising at first sight since we have used only a 2-dimensional smearing. Recall however that in free field theory in Minkowski space, the action of the momentum operator on cylindrical functions is well-defined in the same sense without any smearing at all. In our case, a 2-dimensional smearing is needed because our states contain one rather than three- dimensional excitations. B. Area operators Let us now turn to the integrand of the smeared area operator (corresponding to (3.3)). Denoting the determinant of the intrinsic metric on S by g S, we have: [ĝ S ] f (x) Ψ γ := [E 3 i ] f (x)[e 3i ] f (x) Ψ γ = l4 P 4 [ κ(i, J)f ǫ (x, v αi )f ǫ (x, v αj ) XI i Xi J ] Ψ γ, (3.12) I,J where the summation goes over all the oriented pairs (I, J); v αi and v αj are the vertices at which edges e I and e J intersect S; κ(i, J) = κ I κ J equals 0 if either of the two edges e I and e J fails to intersect S or lies entirely in S, +1 if they lie on the same side of S, and, 1 if they lie on the opposite sides. (For notational simplicity, from now on we shall not keep track of the position of the internal indices i; as noted in Sec. 2.3, they are contracted using the invariant metric on the Lie algebra su(2).) The next step is to consider vertices v α at which γ intersects S and simply rewrite the above sum by re-grouping terms by vertices. The result simplifies if we choose ǫ sufficiently small so that, f ǫ (x, v αi )f ǫ (x, v αj ) is zero unless v αi = v αj. We then have: 5 Given two graphs, γ and γ, we say that γ γ if and only if every edge of γ can be written as a composition of edges of γ. Given two such graphs, there is a projection map from A γ to A γ, which, via pull-back, provides an unitary embedding U γ,γ of H γ o into H γ o. A family of operators O γ on the Hilbert spaces Hγ o is said to be compatible if U γ,γ O γ = O γ U γ,γ and U γ,γ D γ D γ for all γ γ, where D γ and D γ are the domains of O γ and O g. 12

13 [ĝ S ] f (x) Ψ γ = l4 P 4 [ α (f ǫ (x, v α )) 2 I α,j α κ(i α, J α )X i I α X i J α ] Ψ γ, (3.13) where the index α labels the vertices on S and I α and J α label the edges at the vertex α. The next step is to take the square-root of this expression. The same reasoning that established the self-adjointness of [Ê3 i ] f(x) now implies that [ĝ S ] f (x) is a non-negative selfadjoint operator and hence has a well-defined square-root which is also a positive definite self-adjoint operator. Since we have chosen ǫ to be sufficiently small, for any given point x in S, f ǫ (x, v α ) is non-zero for at most one vertex v α. We can therefore take the sum over α outside the square-root. One then obtains ([ĝ S ] f ) 1 2 (x) Ψγ = l2 P 2 f ǫ (x, v α )[ κ(i α, J α )XI i α XJ i α ] 1 2 Ψγ. (3.14) α I α,j α Note that the operator is neatly split; the x-dependence all resides in f ǫ and the operator within the square-root is internal in the sense that it acts only on copies of SU(2). Finally, we can remove the regulator, i.e., take the limit as ǫ tends to zero. By integrating both sides against test functions on S and then taking the limit, we conclude that the following equality holds in the distributional sense: gs (x) Ψ γ = l2 P 2 δ (2) (x, v α )[ κ(i α, J α )XI i α XJ i α ] 1 2 Ψγ. (3.15) α I α,j α Hence, the regularized area operator is given by:  S Ψ γ = l2 P 2 [ κ(i α, J α )XI i α XJ i α ] 1 2 Ψγ. (3.16) α I α,j α (Here, as before, α labels the vertices at which γ intersects S and I α labels the edges of γ at the vertex v α.) With Cyl 2 as its domain,  S is essentially self-adjoint on the Hilbert space H o. Let us now remove the assumption that the surface Σ is covered by a single chart of adapted coordinates. If such a global chart does not exist, we can cover Σ with a family U of neighborhoods such that for each U U there exists a local coordinates system (x a ) adapted to Σ. Let (ϕ U ) U U be a partition of unity associated to U. We just repeat the above regularization for a slightly modified classical surface area functional, namely for A S,U := S dx 1 dx 2 ϕ U [E 3 i E3i ] 1 2 (3.17) which has support within a domain U of an adapted chart. Thus, we obtain the operator ˆ A SU. Then we just define  S = U U  S,U. (3.18) 13

14 The result is given again by the formula (3.16). The reason why the functions ϕ U disappear from the result is that the operator obtained for a single domain of an adapted chart is insensitive on changes of this chart. This concludes our technical discussion. The classical expression A S of (2.2) is a rather complicated. It is therefore somewhat surprising that the corresponding quantum operators can be constructed rigorously and have quite manageable expressions. The essential reason is the underlying diffeomorphism invariance which severely restricts the possible operators. Given a surface and a graph, the only diffeomorphism invariant entities are the intersection vertices. Thus, a diffeomorphism covariant operator can only involve structure at these vertices. In our case, it just acts on the copies of SU(2) associated with various edges at these vertices. We have presented this derivation in considerable detail to spell out all the assumptions, to bring out the generality of the procedure and to illustrate how regularization can be carried out in a fully non-perturbative treatment. While one is free to introduce auxiliary structures such as preferred charts or background fields in the intermediate steps, the final result must respect the underlying diffeomorphism invariance of the theory. These basic ideas will be used repeatedly for other geometric operators in the sequel to this paper. C. General properties of operators 1. Discreteness of the spectrum: By inspection, it follows that the total area operator  S leaves the sub-space of Cyl 2 γ which is associated with any one graph γ invariant and is a self-adjoint operator on the sub-space Hγ o of H o corresponding to γ. Next, recall that Hγ o = L2 (A γ, dµ o ), where A γ is a compact manifold, isomorphic with (SU(2)) N where N is the total number of edges in γ. As, we explained below, the restriction of  S to Hγ o is given by certain commuting elliptic differential operators on this compact manifold. Therefore, all its eigenvalues are discrete. Now suppose that the complete spectrum of  S on H o has a continuous part. Denote by P c the associated projector. Then, given any Ψ in H o, P c Ψ is orthogonal to Hγ o for any graph γ, and hence to the space Cyl of cylindrical functions. Now, since Cyl 2 is dense in H o, P c Ψ must vanish for all Ψ in H o. Hence, the spectrum of ÂS has no continuous part. Note that this method is rather general: It can be used to show that any self-adjoint operator on H o which maps (the intersection of its domain with) Hγ o to Hγ, o and whose action on Hγ o is given by elliptic differential operators, has a purely discrete spectrum on H o. Geometrical operators, constructed purely from the triad field tend to satisfy these properties. 2. Area element: Note that not only is the total area operator well-defined, but in fact it arises from a local area element, g S, which is an operator-valued distribution in the usual sense. Thus, if we integrate it against test functions, the operator is densely defined on H o (with C 2 cylindrical functions as domain) and the matrix elements Ψ γ, g S (x) Ψ γ (3.19) 14

15 are 2-dimensional distributions on S. Furthermore, since we did not have to renormalize the regularized operator (3.14) before removing the regulator, there are no free renormalization constants involved. The local operator is completely unambiguous. 3. [ĝ S ] f versus its square-root: Although the regulated operator [ĝ s ] f is well-defined, if we let ǫ to go zero, the resulting operator is in fact divergent: roughly, it would lead to the square of the 2-dimensional δ distribution. Thus, the determinant of the 2-metric is not a well-defined in the quantum theory. As we saw, however, the square-root of the determinant is well defined: We have to first take the square-root of the regulated expression and then remove the regulator. This, in effect, is the essence of the regularization procedure. To get around this divergence of ĝ S, as is common in Minkowskian field theories, we could have first rescaled [ĝ S ] f ] by an appropriate factor and then taken the limit. Then result can be a well-defined operator, but it will depend on the choice of the regulator, i.e., additional structure introduced in the procedure. Indeed, if the resulting operator is to have the same density character as its classical analog g S (x) which is a scalar density of weight two then the operator can not respect the underlying diffeomorphism invariance. 6 For, there is no metric/chart independent distribution on S of density weight two. Hence, such a renormalized operator is not useful to a fully non-perturbative approach. For the square-root, on the other hand, we need a local density of weight one. And, the 2-dimensional Dirac distribution provides this; now is no apriori obstruction for a satisfactory operator corresponding to the area element to exist. This is an illustration of what appears to be typical in non-perturbative approaches to quantum gravity: Either the limit of the operator exists as the regulator is removed without the need of renormalization or it inherits back-ground dependent renormalization fields (rather than constants). 4. Vertex operators: As noted already, in the final expressions of the area element and area operators, there is a clean separation between the x-dependent and the internal parts. Given a graph γ, the internal part is a sum of square-roots of the operators S,vα := I α,j α κ(i α, J α )X i I α X i J α (3.20) associated with the surface S and the vertex v α on it. It is straightforward to check that operators corresponding to different vertices commute. Therefore, to analyze the properties of area operators, we can focus just on one vertex operator at a time. Furthermore, given the surface S and a point v on it, we can define an operator S,v on the dense sub-space Cyl 2 on H o as follows: 6 If, on the other hand, for some reason, we are willing to allow the limiting operator to have a different density character than its classical analog, one can renormalize [ĝ] f (x) in such a way as to obtain a background independent limit. For instance, we may use f ǫ = (1/ǫ 2 )θ( x x ǫ 2 ), and rescale [ĝ] f by ǫ 2 before taking the limit. Then the limit is a well defined, diffeomorphism covariant operator but it is a scalar density of weight one rather than two. 15

16 { I,J κ(i, J)X S,v Ψ γ := I ixi J Ψ γ, if γ intersects S in v, 0, Otherwise (3.21) where I and J label the edges of γ which have v as a vertex. (Recall that every cylindrical function is associated with some graph γ. As before, if γ intersects S at v but v is not a vertex of γ, one can extend γ just by adding a new vertex v and orienting the edges at v to outgoing.) It is straightforward to verify that this definition is unambiguous: if a cylindrical function can be represented in two different ways, say as Ψ γ and Ψ γ, then S,v Ψ γ and S,v Ψ γ are two representations of the same function on A. There is a precise sense [8] in which S,v can be regarded as a Laplacian operator on H o. The area operator is a sum over all the points v of S of square-roots of Laplacians, Â S = l2 P 2 v S S,v. (3.22) (Here the sum is well defined because, for any cylindrical function, it contains only a finite number of non-zero terms, corresponding to the isolated intersection points of the associated graph with S). We will see in the next subsection that this fact is reflected in its spectrum. 5. Gauge invariance: The classical area element g S is invariant under the internal rotations of triads Ei a ; its Poisson bracket with the Gauss constraint functional vanishes. This symmetry is preserved in the quantum theory: the quantum operator g S commutes with the induced action of G on the Hilbert space H o. Thus, g S and the total area operator ÂS map the space of gauge invariant states to itself; they project down to the Hilbert space H o of kinematic states. Note, however, that the regulated triad operators [Ê3 i ] f are not gauge invariant; they are defined only on H o. Nonetheless, they are useful; they feature in an important way in our regularization scheme. In the loop representation, by contrast, one can only introduce gauge invariant operators and hence the regulated triad operators do not exist. Furthermore, even in the definition (3.3) of the regularized area element, one must use holonomies to transport indices between the two points y and z. While this manifest gauge invariance is conceptually pleasing, in practice it often makes the calculations in the loop representation cumbersome; one has to keep track of these holonomy insertions in the intermediate steps although they do not contribute to the final result. 6. Overall Factors: The overall numerical factors in the expressions of various operators considered above depend on two conventions. The first is the convention noted in footnote 4 used in the regularization procedure. Could we not have used a different convention, setting 0 dzg(z)δ(z) = cg(0) and 0 dzg(z)δ(z) = (1 c)g(0) for some constant c 1/2? The answer is in the negative. For, in this case, the constant κ I would take values: 0, if e I is tangential to S or does not intersect S, κ I = +2c, if e I has an isolated intersection with S and lies above S (3.23) 2(1 c), if e I has an isolated intersection with S and lies below S 16

17 It then follows that, unless c = 1/2, the action of the area operator ÂS on a given cylindrical function would change if we simply reverse the orientation on S (keeping the orientation on Σ the same). Since this is physically inadmissible, we must have c = 1/2; there is really no freedom in this part of the regularization procedure. The second convention has to do with the overall numerical factor in the action, which dictates the numerical coefficients in the symplectic structure. Here, we have adopted the convention of [25] (see chapter 9) which makes the Poisson bracket {A i a(x), Ej(y)} b = Gδa bδi j δ(x, y), enabling us to express Êa i (x) as ig hδ/δai a (x). (Had we rescaled the action by 1/8π as is sometimes done, in our expressions, Newton s constant G would be replaced by 8πG.) IV. EIGENVALUES AND EIGENVECTORS This section is divided into three parts. In the first, we derive the complete spectrum of the area operators, in the second, we extend the notion of spin networks, and in the third, we use this extension to discuss eigenvectors. A. The complete spectrum We are now ready to calculate the complete spectrum of  S. Since ÂS is a sum of square-roots of vertex operators which all commute with one another, the task reduces to that of finding the spectrum of each vertex operator. Furthermore, since vertex operators map (C 2 ) cylindrical functions associated with any one graph to (C 0 ) cylindrical functions associated with the same graph, we can begin with an arbitrary but fixed graph γ. Consider then a vertex operator S,v and focus on the edges of γ which intersect S at v. Let us divide the edges in to three categories: let e 1,..., e d lie below S ( down ), e d+1,..., e u lie above S ( up ) and let e u+1,..., e t be tangential to S. (As before, the labels down and up do not have an invariant significance; the orientation of S and of Σ enable us to divide the non-tangential edges in to two parts and we just label one as down and the other as up.) Let us set: J (d) S,v J (t) S,v iγ := i (X i Xi d iγ := i (X i u Xi t i ), J(u) S,v ), J(d+u) S,v γ := i (Xd+1 i Xi u ), iγ:= J (d) i (u) i S,v + J S,v (4.1) where XI i is the operator defined in (3.11) assigned to the point v and an edge e I at v. This notation is suggestive. We can associate with each edge e a particle with only a spin degree of freedom. Then, the operators ixe i can be thought of as the ith component i and J (t) S,vi as of angular momentum operators associated with that particle and J (d) i (u) S,v, J S,v the total down, up and tangential angular momentum operators at the vertex v. By varying the graph, we thus obtain a family of operators. It is easy to check that they satisfy the compatibility conditions and thus define operators J (u) i (d) i (t) S,v, J S,v.J S,vi and 17

18 J (d+u) i S,v on Cyl. It is also easy to verify that they all commute with one another. Hence, one can express the vertex operator S,v simply as: S,v = (J (d) S,v i J (u) S,v i (d) i (u) i )(J S,v J S,v ) ; (4.2) because of the factor κ(i, J) in (3.21), the edges which are tangential do not feature in this expression. The evaluation of possible eigenvalues is now straightforward. It is simplest to express S,v as S,v = 2(J (d) S,v) 2 + 2(J (u) S,v) 2 (J (d+u) S,v ) 2. (4.3) and, as in elementary text-books, go to the representation in which the operators (J S,v) (d) 2, (J S,v) (u) 2 and (J (d+u) S,v ) 2 are diagonal. If we restrict now the operators to Cyl γ associated to a fixed graph, it is obvious that the possible eigenvalues λ of S,v are given by: λ S,v = 2j (d) (j (d) + 1) + 2j (u) (j (u) + 1) j (d+u) (j (d+u) + 1) (4.4) where j (d), j (u) and j (d+u) are half integers subject to the usual condition: j (d+u) { j (d) j (u), j (d) j (u) + 1,..., j (d) + j (u) }. (4.5) Returning to the total area operator, we note that the vertex operators associated with distinct vertices commute. Although the sum (3.22) is not finite, restricted to any graph γ and Cyl γ it becomes finite. Therefore, the eigenvalues a S of ÂS are given by a S = l2 P 2 α [ 2j (d) α (j(d) α + 1) + 2j(u) α (j(u) α + 1) j(d+u) α (j (d+u) α + 1) ]1 2 (4.6) where α labels a finite set of points in S and the non-negative half-integers assigned to each α are subject to the inequality (4.5). The question now is if all these eigenvalues are actually attained, i.e., if, given any a S of the form (4.6), there are eigenvectors in H o with that eigenvalue. In Sec. 4.3, will show that the full spectrum is indeed realized on H o. The area operators map the subspace H o of gauge invariant elements of H o to itself. Hence we can ask for their spectrum on H o. We will see in Sec. 4.3 that further restrictions can now arise depending on the topology of the surface S. There are three cases: i) The case when S is an open surface whose closure is contained in Σ. An example is provided by the disk z = 0, x 2 +y 2 < r o in R 3. In this case, there is no additional condition; all a S of (4.6) subject to (4.5) are realized. ii) The case when the surface S is closed ( S = ) and divides Σ into disjoint open sets Σ 1 and Σ 2 (i.e., Σ = Σ 1 S Σ 2 with Σ 1 Σ 2 =.) An example is given by: 18

arxiv:gr-qc/ v2 12 Aug 1998

arxiv:gr-qc/ v2 12 Aug 1998 Quantum Theory of Geometry III: Non-commutativity of Riemannian Structures arxiv:gr-qc/9806041v2 12 Aug 1998 Abhay Ashtekar 1,3, Alejandro Corichi 1,2, and José A. Zapata 1 1 Center for Gravitational Physics

More information

Irreducibility of the Ashtekar Isham Lewandowski Representation

Irreducibility of the Ashtekar Isham Lewandowski Representation Irreducibility of the Ashtekar Isham Lewandowski Representation Hanno Sahlmann, Center for Gravitational Physics and Geometry, The Pennsylvania State University, University Park, PA, USA Thomas Thiemann,

More information

BRST and Dirac Cohomology

BRST and Dirac Cohomology BRST and Dirac Cohomology Peter Woit Columbia University Dartmouth Math Dept., October 23, 2008 Peter Woit (Columbia University) BRST and Dirac Cohomology October 2008 1 / 23 Outline 1 Introduction 2 Representation

More information

CALCULUS ON MANIFOLDS. 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M =

CALCULUS ON MANIFOLDS. 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M = CALCULUS ON MANIFOLDS 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M = a M T am, called the tangent bundle, is itself a smooth manifold, dim T M = 2n. Example 1.

More information

Introduction to Group Theory

Introduction to Group Theory Chapter 10 Introduction to Group Theory Since symmetries described by groups play such an important role in modern physics, we will take a little time to introduce the basic structure (as seen by a physicist)

More information

Mathematical Introduction

Mathematical Introduction Chapter 1 Mathematical Introduction HW #1: 164, 165, 166, 181, 182, 183, 1811, 1812, 114 11 Linear Vector Spaces: Basics 111 Field A collection F of elements a,b etc (also called numbers or scalars) with

More information

QM and Angular Momentum

QM and Angular Momentum Chapter 5 QM and Angular Momentum 5. Angular Momentum Operators In your Introductory Quantum Mechanics (QM) course you learned about the basic properties of low spin systems. Here we want to review that

More information

Diffeomorphism covariant representations of the holonomy-flux -algebra arxiv:gr-qc/ v1 14 Feb 2003

Diffeomorphism covariant representations of the holonomy-flux -algebra arxiv:gr-qc/ v1 14 Feb 2003 Diffeomorphism covariant representations of the holonomy-flux -algebra arxiv:gr-qc/0302059v1 14 Feb 2003 Andrzej Oko lów 1 and Jerzy Lewandowski 2,1,3 February 14, 2003 1. Instytut Fizyki Teoretycznej,

More information

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom Representation theory and quantum mechanics tutorial Spin and the hydrogen atom Justin Campbell August 3, 2017 1 Representations of SU 2 and SO 3 (R) 1.1 The following observation is long overdue. Proposition

More information

Lecture 10: A (Brief) Introduction to Group Theory (See Chapter 3.13 in Boas, 3rd Edition)

Lecture 10: A (Brief) Introduction to Group Theory (See Chapter 3.13 in Boas, 3rd Edition) Lecture 0: A (Brief) Introduction to Group heory (See Chapter 3.3 in Boas, 3rd Edition) Having gained some new experience with matrices, which provide us with representations of groups, and because symmetries

More information

The following definition is fundamental.

The following definition is fundamental. 1. Some Basics from Linear Algebra With these notes, I will try and clarify certain topics that I only quickly mention in class. First and foremost, I will assume that you are familiar with many basic

More information

Incompatibility Paradoxes

Incompatibility Paradoxes Chapter 22 Incompatibility Paradoxes 22.1 Simultaneous Values There is never any difficulty in supposing that a classical mechanical system possesses, at a particular instant of time, precise values of

More information

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory Part V 7 Introduction: What are measures and why measurable sets Lebesgue Integration Theory Definition 7. (Preliminary). A measure on a set is a function :2 [ ] such that. () = 2. If { } = is a finite

More information

1 Mathematical preliminaries

1 Mathematical preliminaries 1 Mathematical preliminaries The mathematical language of quantum mechanics is that of vector spaces and linear algebra. In this preliminary section, we will collect the various definitions and mathematical

More information

Loop Quantum Gravity 2. The quantization : discreteness of space

Loop Quantum Gravity 2. The quantization : discreteness of space Loop Quantum Gravity 2. The quantization : discreteness of space Karim NOUI Laboratoire de Mathématiques et de Physique Théorique, TOURS Astro Particules et Cosmologie, PARIS Clermont - Ferrand ; january

More information

Gauge Fixing and Constrained Dynamics in Numerical Relativity

Gauge Fixing and Constrained Dynamics in Numerical Relativity Gauge Fixing and Constrained Dynamics in Numerical Relativity Jon Allen The Dirac formalism for dealing with constraints in a canonical Hamiltonian formulation is reviewed. Gauge freedom is discussed and

More information

GEOMETRIC QUANTIZATION

GEOMETRIC QUANTIZATION GEOMETRIC QUANTIZATION 1. The basic idea The setting of the Hamiltonian version of classical (Newtonian) mechanics is the phase space (position and momentum), which is a symplectic manifold. The typical

More information

Symmetries, Groups, and Conservation Laws

Symmetries, Groups, and Conservation Laws Chapter Symmetries, Groups, and Conservation Laws The dynamical properties and interactions of a system of particles and fields are derived from the principle of least action, where the action is a 4-dimensional

More information

Connection Variables in General Relativity

Connection Variables in General Relativity Connection Variables in General Relativity Mauricio Bustamante Londoño Instituto de Matemáticas UNAM Morelia 28/06/2008 Mauricio Bustamante Londoño (UNAM) Connection Variables in General Relativity 28/06/2008

More information

Panel Discussion on: Recovering Low Energy Physics

Panel Discussion on: Recovering Low Energy Physics p. Panel Discussion on: Recovering Low Energy Physics Abhay Ashtekar, Laurent Freidel, Carlo Rovelli Continuation of the discussion on semi-classical issues from two weeks ago following John Barret s talk.

More information

The Spinor Representation

The Spinor Representation The Spinor Representation Math G4344, Spring 2012 As we have seen, the groups Spin(n) have a representation on R n given by identifying v R n as an element of the Clifford algebra C(n) and having g Spin(n)

More information

Bianchi I Space-times and Loop Quantum Cosmology

Bianchi I Space-times and Loop Quantum Cosmology Bianchi I Space-times and Loop Quantum Cosmology Edward Wilson-Ewing Institute for Gravitation and the Cosmos The Pennsylvania State University Work with Abhay Ashtekar October 23, 2008 E. Wilson-Ewing

More information

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f))

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f)) 1. Basic algebra of vector fields Let V be a finite dimensional vector space over R. Recall that V = {L : V R} is defined to be the set of all linear maps to R. V is isomorphic to V, but there is no canonical

More information

Generators for Continuous Coordinate Transformations

Generators for Continuous Coordinate Transformations Page 636 Lecture 37: Coordinate Transformations: Continuous Passive Coordinate Transformations Active Coordinate Transformations Date Revised: 2009/01/28 Date Given: 2009/01/26 Generators for Continuous

More information

INSTANTON MODULI AND COMPACTIFICATION MATTHEW MAHOWALD

INSTANTON MODULI AND COMPACTIFICATION MATTHEW MAHOWALD INSTANTON MODULI AND COMPACTIFICATION MATTHEW MAHOWALD () Instanton (definition) (2) ADHM construction (3) Compactification. Instantons.. Notation. Throughout this talk, we will use the following notation:

More information

LECTURE 3: Quantization and QFT

LECTURE 3: Quantization and QFT LECTURE 3: Quantization and QFT Robert Oeckl IQG-FAU & CCM-UNAM IQG FAU Erlangen-Nürnberg 14 November 2013 Outline 1 Classical field theory 2 Schrödinger-Feynman quantization 3 Klein-Gordon Theory Classical

More information

Attempts at relativistic QM

Attempts at relativistic QM Attempts at relativistic QM based on S-1 A proper description of particle physics should incorporate both quantum mechanics and special relativity. However historically combining quantum mechanics and

More information

Polymer Parametrized field theory

Polymer Parametrized field theory .... Polymer Parametrized field theory Alok Laddha Raman Research Institute, India Dec 1 Collaboration with : Madhavan Varadarajan (Raman Research Institute, India) Alok Laddha (RRI) Polymer Parametrized

More information

1 Math 241A-B Homework Problem List for F2015 and W2016

1 Math 241A-B Homework Problem List for F2015 and W2016 1 Math 241A-B Homework Problem List for F2015 W2016 1.1 Homework 1. Due Wednesday, October 7, 2015 Notation 1.1 Let U be any set, g be a positive function on U, Y be a normed space. For any f : U Y let

More information

A Note On The Chern-Simons And Kodama Wavefunctions

A Note On The Chern-Simons And Kodama Wavefunctions hep-th/0306083 arxiv:gr-qc/0306083v2 19 Jun 2003 A Note On The Chern-Simons And Kodama Wavefunctions Edward Witten Institute For Advanced Study, Princeton NJ 08540 USA Yang-Mills theory in four dimensions

More information

Intrinsic Time Quantum Geometrodynamics (ITQG)

Intrinsic Time Quantum Geometrodynamics (ITQG) Intrinsic Time Quantum Geometrodynamics (ITQG) Assistant Professor Eyo Ita Eyo Eyo Ita Physics Department LQG International Seminar United States Naval Academy Annapolis, MD 27 October, 2015 Outline of

More information

2. Lie groups as manifolds. SU(2) and the three-sphere. * version 1.4 *

2. Lie groups as manifolds. SU(2) and the three-sphere. * version 1.4 * . Lie groups as manifolds. SU() and the three-sphere. * version 1.4 * Matthew Foster September 1, 017 Contents.1 The Haar measure 1. The group manifold for SU(): S 3 3.3 Left- and right- group translations

More information

Donoghue, Golowich, Holstein Chapter 4, 6

Donoghue, Golowich, Holstein Chapter 4, 6 1 Week 7: Non linear sigma models and pion lagrangians Reading material from the books Burgess-Moore, Chapter 9.3 Donoghue, Golowich, Holstein Chapter 4, 6 Weinberg, Chap. 19 1 Goldstone boson lagrangians

More information

arxiv:gr-qc/ v1 2 Nov 1994

arxiv:gr-qc/ v1 2 Nov 1994 Spin Network States in Gauge Theory John C. Baez Department of Mathematics University of California Riverside CA 92521 November 1, 1994 arxiv:gr-qc/9411007v1 2 Nov 1994 Abstract Given a real-analytic manifold

More information

Manifestly diffeomorphism invariant classical Exact Renormalization Group

Manifestly diffeomorphism invariant classical Exact Renormalization Group Manifestly diffeomorphism invariant classical Exact Renormalization Group Anthony W. H. Preston University of Southampton Supervised by Prof. Tim R. Morris Talk prepared for Asymptotic Safety seminar,

More information

Mathematical structure of loop quantum cosmology

Mathematical structure of loop quantum cosmology c 2003 International Press Adv. Theor. Math. Phys. 7 (2003) 233 268 arxiv:gr-qc/0304074v4 24 Dec 2003 Mathematical structure of loop quantum cosmology Abhay Ashtekar 1,2, Martin Bojowald 1,2, and Jerzy

More information

Approaches to Quantum Gravity A conceptual overview

Approaches to Quantum Gravity A conceptual overview Approaches to Quantum Gravity A conceptual overview Robert Oeckl Instituto de Matemáticas UNAM, Morelia Centro de Radioastronomía y Astrofísica UNAM, Morelia 14 February 2008 Outline 1 Introduction 2 Different

More information

Loop Quantum Gravity. Carlo Rovelli. String 08 Genève, August 2008

Loop Quantum Gravity. Carlo Rovelli. String 08 Genève, August 2008 Loop Quantum Gravity Carlo Rovelli String 08 Genève, August 2008 1 i. Loop Quantum Gravity - The problem addressed - Why loops? ii. The theory - Kinematics: spin networks and quanta of space - Dynamics:

More information

Chapter 2 The Group U(1) and its Representations

Chapter 2 The Group U(1) and its Representations Chapter 2 The Group U(1) and its Representations The simplest example of a Lie group is the group of rotations of the plane, with elements parametrized by a single number, the angle of rotation θ. It is

More information

MP463 QUANTUM MECHANICS

MP463 QUANTUM MECHANICS MP463 QUANTUM MECHANICS Introduction Quantum theory of angular momentum Quantum theory of a particle in a central potential - Hydrogen atom - Three-dimensional isotropic harmonic oscillator (a model of

More information

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms (February 24, 2017) 08a. Operators on Hilbert spaces Paul Garrett garrett@math.umn.edu http://www.math.umn.edu/ garrett/ [This document is http://www.math.umn.edu/ garrett/m/real/notes 2016-17/08a-ops

More information

3 Quantization of the Dirac equation

3 Quantization of the Dirac equation 3 Quantization of the Dirac equation 3.1 Identical particles As is well known, quantum mechanics implies that no measurement can be performed to distinguish particles in the same quantum state. Elementary

More information

Clifford Algebras and Spin Groups

Clifford Algebras and Spin Groups Clifford Algebras and Spin Groups Math G4344, Spring 2012 We ll now turn from the general theory to examine a specific class class of groups: the orthogonal groups. Recall that O(n, R) is the group of

More information

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces. Math 350 Fall 2011 Notes about inner product spaces In this notes we state and prove some important properties of inner product spaces. First, recall the dot product on R n : if x, y R n, say x = (x 1,...,

More information

Select/ Special Topics in Atomic Physics Prof. P. C. Deshmukh Department of Physics Indian Institute of Technology, Madras

Select/ Special Topics in Atomic Physics Prof. P. C. Deshmukh Department of Physics Indian Institute of Technology, Madras Select/ Special Topics in Atomic Physics Prof. P. C. Deshmukh Department of Physics Indian Institute of Technology, Madras Lecture No. # 06 Angular Momentum in Quantum Mechanics Greetings, we will begin

More information

Lecture I: Constrained Hamiltonian systems

Lecture I: Constrained Hamiltonian systems Lecture I: Constrained Hamiltonian systems (Courses in canonical gravity) Yaser Tavakoli December 15, 2014 1 Introduction In canonical formulation of general relativity, geometry of space-time is given

More information

On a Derivation of the Dirac Hamiltonian From a Construction of Quantum Gravity

On a Derivation of the Dirac Hamiltonian From a Construction of Quantum Gravity On a Derivation of the Dirac Hamiltonian From a Construction of Quantum Gravity Johannes Aastrup a1, Jesper Møller Grimstrup b2 & Mario Paschke c3 arxiv:1003.3802v1 [hep-th] 19 Mar 2010 a Mathematisches

More information

1 Quantum fields in Minkowski spacetime

1 Quantum fields in Minkowski spacetime 1 Quantum fields in Minkowski spacetime The theory of quantum fields in curved spacetime is a generalization of the well-established theory of quantum fields in Minkowski spacetime. To a great extent,

More information

arxiv:gr-qc/ v2 17 Nov 1993

arxiv:gr-qc/ v2 17 Nov 1993 SU-GP-93/7-6 CGPG-93/8-3 gr-qc/930808 Solution space of 2+ gravity on R T 2 in Witten s connection formulation arxiv:gr-qc/930808v2 7 Nov 993 Jorma Louko and Donald M. Marolf Department of Physics, Syracuse

More information

arxiv: v1 [gr-qc] 5 Apr 2014

arxiv: v1 [gr-qc] 5 Apr 2014 Quantum Holonomy Theory Johannes Aastrup a1 & Jesper Møller Grimstrup 2 a Mathematisches Institut, Universität Hannover, Welfengarten 1, D-30167 Hannover, Germany. arxiv:1404.1500v1 [gr-qc] 5 Apr 2014

More information

Geometry in a Fréchet Context: A Projective Limit Approach

Geometry in a Fréchet Context: A Projective Limit Approach Geometry in a Fréchet Context: A Projective Limit Approach Geometry in a Fréchet Context: A Projective Limit Approach by C.T.J. Dodson University of Manchester, Manchester, UK George Galanis Hellenic

More information

,, rectilinear,, spherical,, cylindrical. (6.1)

,, rectilinear,, spherical,, cylindrical. (6.1) Lecture 6 Review of Vectors Physics in more than one dimension (See Chapter 3 in Boas, but we try to take a more general approach and in a slightly different order) Recall that in the previous two lectures

More information

arxiv:gr-qc/ v1 31 Jul 2001

arxiv:gr-qc/ v1 31 Jul 2001 SINGULARITY AVOIDANCE BY COLLAPSING SHELLS IN QUANTUM GRAVITY 1 arxiv:gr-qc/0107102v1 31 Jul 2001 Petr Hájíček Institut für Theoretische Physik, Universität Bern, Sidlerstrasse 5, 3012 Bern, Switzerland.

More information

2-Form Gravity of the Lorentzian Signature

2-Form Gravity of the Lorentzian Signature 2-Form Gravity of the Lorentzian Signature Jerzy Lewandowski 1 and Andrzej Oko lów 2 Instytut Fizyki Teoretycznej, Uniwersytet Warszawski, ul. Hoża 69, 00-681 Warszawa, Poland arxiv:gr-qc/9911121v1 30

More information

The quantum state as a vector

The quantum state as a vector The quantum state as a vector February 6, 27 Wave mechanics In our review of the development of wave mechanics, we have established several basic properties of the quantum description of nature:. A particle

More information

1 Differentiable manifolds and smooth maps

1 Differentiable manifolds and smooth maps 1 Differentiable manifolds and smooth maps Last updated: April 14, 2011. 1.1 Examples and definitions Roughly, manifolds are sets where one can introduce coordinates. An n-dimensional manifold is a set

More information

Consistent Histories. Chapter Chain Operators and Weights

Consistent Histories. Chapter Chain Operators and Weights Chapter 10 Consistent Histories 10.1 Chain Operators and Weights The previous chapter showed how the Born rule can be used to assign probabilities to a sample space of histories based upon an initial state

More information

Recitation 1 (Sep. 15, 2017)

Recitation 1 (Sep. 15, 2017) Lecture 1 8.321 Quantum Theory I, Fall 2017 1 Recitation 1 (Sep. 15, 2017) 1.1 Simultaneous Diagonalization In the last lecture, we discussed the situations in which two operators can be simultaneously

More information

Quantum Theory and Group Representations

Quantum Theory and Group Representations Quantum Theory and Group Representations Peter Woit Columbia University LaGuardia Community College, November 1, 2017 Queensborough Community College, November 15, 2017 Peter Woit (Columbia University)

More information

Cambridge University Press The Mathematics of Signal Processing Steven B. Damelin and Willard Miller Excerpt More information

Cambridge University Press The Mathematics of Signal Processing Steven B. Damelin and Willard Miller Excerpt More information Introduction Consider a linear system y = Φx where Φ can be taken as an m n matrix acting on Euclidean space or more generally, a linear operator on a Hilbert space. We call the vector x a signal or input,

More information

Vector Spaces. Vector space, ν, over the field of complex numbers, C, is a set of elements a, b,..., satisfying the following axioms.

Vector Spaces. Vector space, ν, over the field of complex numbers, C, is a set of elements a, b,..., satisfying the following axioms. Vector Spaces Vector space, ν, over the field of complex numbers, C, is a set of elements a, b,..., satisfying the following axioms. For each two vectors a, b ν there exists a summation procedure: a +

More information

The Hurewicz Theorem

The Hurewicz Theorem The Hurewicz Theorem April 5, 011 1 Introduction The fundamental group and homology groups both give extremely useful information, particularly about path-connected spaces. Both can be considered as functors,

More information

Hilbert Spaces. Hilbert space is a vector space with some extra structure. We start with formal (axiomatic) definition of a vector space.

Hilbert Spaces. Hilbert space is a vector space with some extra structure. We start with formal (axiomatic) definition of a vector space. Hilbert Spaces Hilbert space is a vector space with some extra structure. We start with formal (axiomatic) definition of a vector space. Vector Space. Vector space, ν, over the field of complex numbers,

More information

MAT265 Mathematical Quantum Mechanics Brief Review of the Representations of SU(2)

MAT265 Mathematical Quantum Mechanics Brief Review of the Representations of SU(2) MAT65 Mathematical Quantum Mechanics Brief Review of the Representations of SU() (Notes for MAT80 taken by Shannon Starr, October 000) There are many references for representation theory in general, and

More information

Vectors. January 13, 2013

Vectors. January 13, 2013 Vectors January 13, 2013 The simplest tensors are scalars, which are the measurable quantities of a theory, left invariant by symmetry transformations. By far the most common non-scalars are the vectors,

More information

Loop Quantum Gravity a general-covariant lattice gauge theory. Francesca Vidotto UNIVERSITY OF THE BASQUE COUNTRY

Loop Quantum Gravity a general-covariant lattice gauge theory. Francesca Vidotto UNIVERSITY OF THE BASQUE COUNTRY a general-covariant lattice gauge theory UNIVERSITY OF THE BASQUE COUNTRY Bad Honnef - August 2 nd, 2018 THE GRAVITATIONAL FIELD GENERAL RELATIVITY: background independence! U(1) SU(2) SU(3) SL(2,C) l

More information

Towards a manifestly diffeomorphism invariant Exact Renormalization Group

Towards a manifestly diffeomorphism invariant Exact Renormalization Group Towards a manifestly diffeomorphism invariant Exact Renormalization Group Anthony W. H. Preston University of Southampton Supervised by Prof. Tim R. Morris Talk prepared for UK QFT-V, University of Nottingham,

More information

Generalized Global Symmetries

Generalized Global Symmetries Generalized Global Symmetries Anton Kapustin Simons Center for Geometry and Physics, Stony Brook April 9, 2015 Anton Kapustin (Simons Center for Geometry and Physics, Generalized StonyGlobal Brook) Symmetries

More information

II. DIFFERENTIABLE MANIFOLDS. Washington Mio CENTER FOR APPLIED VISION AND IMAGING SCIENCES

II. DIFFERENTIABLE MANIFOLDS. Washington Mio CENTER FOR APPLIED VISION AND IMAGING SCIENCES II. DIFFERENTIABLE MANIFOLDS Washington Mio Anuj Srivastava and Xiuwen Liu (Illustrations by D. Badlyans) CENTER FOR APPLIED VISION AND IMAGING SCIENCES Florida State University WHY MANIFOLDS? Non-linearity

More information

Page 52. Lecture 3: Inner Product Spaces Dual Spaces, Dirac Notation, and Adjoints Date Revised: 2008/10/03 Date Given: 2008/10/03

Page 52. Lecture 3: Inner Product Spaces Dual Spaces, Dirac Notation, and Adjoints Date Revised: 2008/10/03 Date Given: 2008/10/03 Page 5 Lecture : Inner Product Spaces Dual Spaces, Dirac Notation, and Adjoints Date Revised: 008/10/0 Date Given: 008/10/0 Inner Product Spaces: Definitions Section. Mathematical Preliminaries: Inner

More information

Lorentz-covariant spectrum of single-particle states and their field theory Physics 230A, Spring 2007, Hitoshi Murayama

Lorentz-covariant spectrum of single-particle states and their field theory Physics 230A, Spring 2007, Hitoshi Murayama Lorentz-covariant spectrum of single-particle states and their field theory Physics 30A, Spring 007, Hitoshi Murayama 1 Poincaré Symmetry In order to understand the number of degrees of freedom we need

More information

Chapter 2. Linear Algebra. rather simple and learning them will eventually allow us to explain the strange results of

Chapter 2. Linear Algebra. rather simple and learning them will eventually allow us to explain the strange results of Chapter 2 Linear Algebra In this chapter, we study the formal structure that provides the background for quantum mechanics. The basic ideas of the mathematical machinery, linear algebra, are rather simple

More information

Math 676. A compactness theorem for the idele group. and by the product formula it lies in the kernel (A K )1 of the continuous idelic norm

Math 676. A compactness theorem for the idele group. and by the product formula it lies in the kernel (A K )1 of the continuous idelic norm Math 676. A compactness theorem for the idele group 1. Introduction Let K be a global field, so K is naturally a discrete subgroup of the idele group A K and by the product formula it lies in the kernel

More information

Coordinate free non abelian geometry I: the quantum case of simplicial manifolds.

Coordinate free non abelian geometry I: the quantum case of simplicial manifolds. Coordinate free non abelian geometry I: the quantum case of simplicial manifolds. Johan Noldus May 6, 07 Abstract We study the geometry of a simplicial complexes from an algebraic point of view and devise

More information

Tr (Q(E)P (F )) = µ(e)µ(f ) (1)

Tr (Q(E)P (F )) = µ(e)µ(f ) (1) Some remarks on a problem of Accardi G. CASSINELLI and V. S. VARADARAJAN 1. Introduction Let Q, P be the projection valued measures on the real line R that are associated to the position and momentum observables

More information

Stochastic Histories. Chapter Introduction

Stochastic Histories. Chapter Introduction Chapter 8 Stochastic Histories 8.1 Introduction Despite the fact that classical mechanics employs deterministic dynamical laws, random dynamical processes often arise in classical physics, as well as in

More information

As always, the story begins with Riemann surfaces or just (real) surfaces. (As we have already noted, these are nearly the same thing).

As always, the story begins with Riemann surfaces or just (real) surfaces. (As we have already noted, these are nearly the same thing). An Interlude on Curvature and Hermitian Yang Mills As always, the story begins with Riemann surfaces or just (real) surfaces. (As we have already noted, these are nearly the same thing). Suppose we wanted

More information

Quantum Mechanics without Spacetime II

Quantum Mechanics without Spacetime II Quantum Mechanics without Spacetime II - Noncommutative geometry and the free point particle - T. P. Singh Tata Institute of Fundamental Research Homi Bhabha Road, Mumbai 400 005, India email address:

More information

(1.1) In particular, ψ( q 1, m 1 ; ; q N, m N ) 2 is the probability to find the first particle

(1.1) In particular, ψ( q 1, m 1 ; ; q N, m N ) 2 is the probability to find the first particle Chapter 1 Identical particles 1.1 Distinguishable particles The Hilbert space of N has to be a subspace H = N n=1h n. Observables Ân of the n-th particle are self-adjoint operators of the form 1 1 1 1

More information

Eilenberg-Steenrod properties. (Hatcher, 2.1, 2.3, 3.1; Conlon, 2.6, 8.1, )

Eilenberg-Steenrod properties. (Hatcher, 2.1, 2.3, 3.1; Conlon, 2.6, 8.1, ) II.3 : Eilenberg-Steenrod properties (Hatcher, 2.1, 2.3, 3.1; Conlon, 2.6, 8.1, 8.3 8.5 Definition. Let U be an open subset of R n for some n. The de Rham cohomology groups (U are the cohomology groups

More information

Exercises in Geometry II University of Bonn, Summer semester 2015 Professor: Prof. Christian Blohmann Assistant: Saskia Voss Sheet 1

Exercises in Geometry II University of Bonn, Summer semester 2015 Professor: Prof. Christian Blohmann Assistant: Saskia Voss Sheet 1 Assistant: Saskia Voss Sheet 1 1. Conformal change of Riemannian metrics [3 points] Let (M, g) be a Riemannian manifold. A conformal change is a nonnegative function λ : M (0, ). Such a function defines

More information

Solutions to the Hamilton-Jacobi equation as Lagrangian submanifolds

Solutions to the Hamilton-Jacobi equation as Lagrangian submanifolds Solutions to the Hamilton-Jacobi equation as Lagrangian submanifolds Matias Dahl January 2004 1 Introduction In this essay we shall study the following problem: Suppose is a smooth -manifold, is a function,

More information

General Fourier Analysis and Gravitational Waves.

General Fourier Analysis and Gravitational Waves. General Fourier Analysis and Gravitational Waves. Johan Noldus April 0, 06 Abstract In a recent paper [] of this author, we generalized quantum field theory to any curved spacetime. The key idea of the

More information

Density Matrices. Chapter Introduction

Density Matrices. Chapter Introduction Chapter 15 Density Matrices 15.1 Introduction Density matrices are employed in quantum mechanics to give a partial description of a quantum system, one from which certain details have been omitted. For

More information

PLEASE LET ME KNOW IF YOU FIND TYPOS (send to

PLEASE LET ME KNOW IF YOU FIND TYPOS (send  to Teoretisk Fysik KTH Advanced QM (SI2380), Lecture 2 (Summary of concepts) 1 PLEASE LET ME KNOW IF YOU FIND TYPOS (send email to langmann@kth.se) The laws of QM 1. I now discuss the laws of QM and their

More information

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19 Page 404 Lecture : Simple Harmonic Oscillator: Energy Basis Date Given: 008/11/19 Date Revised: 008/11/19 Coordinate Basis Section 6. The One-Dimensional Simple Harmonic Oscillator: Coordinate Basis Page

More information

Physics 70007, Fall 2009 Answers to Final Exam

Physics 70007, Fall 2009 Answers to Final Exam Physics 70007, Fall 009 Answers to Final Exam December 17, 009 1. Quantum mechanical pictures a Demonstrate that if the commutation relation [A, B] ic is valid in any of the three Schrodinger, Heisenberg,

More information

Introduction. Chapter The Purpose of Statistical Mechanics

Introduction. Chapter The Purpose of Statistical Mechanics Chapter 1 Introduction 1.1 The Purpose of Statistical Mechanics Statistical Mechanics is the mechanics developed to treat a collection of a large number of atoms or particles. Such a collection is, for

More information

Observables in the GBF

Observables in the GBF Observables in the GBF Max Dohse Centro de Ciencias Matematicas (CCM) UNAM Campus Morelia M. Dohse (CCM-UNAM Morelia) GBF: Observables GBF Seminar (14.Mar.2013) 1 / 36 References:. [RO2011] R. Oeckl: Observables

More information

Topics in Representation Theory: Roots and Complex Structures

Topics in Representation Theory: Roots and Complex Structures Topics in Representation Theory: Roots and Complex Structures 1 More About Roots To recap our story so far: we began by identifying an important abelian subgroup of G, the maximal torus T. By restriction

More information

Quantization of scalar fields

Quantization of scalar fields Quantization of scalar fields March 8, 06 We have introduced several distinct types of fields, with actions that give their field equations. These include scalar fields, S α ϕ α ϕ m ϕ d 4 x and complex

More information

Geometry and Physics. Amer Iqbal. March 4, 2010

Geometry and Physics. Amer Iqbal. March 4, 2010 March 4, 2010 Many uses of Mathematics in Physics The language of the physical world is mathematics. Quantitative understanding of the world around us requires the precise language of mathematics. Symmetries

More information

Noncommutative geometry and quantum field theory

Noncommutative geometry and quantum field theory Noncommutative geometry and quantum field theory Graeme Segal The beginning of noncommutative geometry is the observation that there is a rough equivalence contravariant between the category of topological

More information

An introduction to Birkhoff normal form

An introduction to Birkhoff normal form An introduction to Birkhoff normal form Dario Bambusi Dipartimento di Matematica, Universitá di Milano via Saldini 50, 0133 Milano (Italy) 19.11.14 1 Introduction The aim of this note is to present an

More information

Math 302 Outcome Statements Winter 2013

Math 302 Outcome Statements Winter 2013 Math 302 Outcome Statements Winter 2013 1 Rectangular Space Coordinates; Vectors in the Three-Dimensional Space (a) Cartesian coordinates of a point (b) sphere (c) symmetry about a point, a line, and a

More information

The Effective Equations for Bianchi IX Loop Quantum Cosmology

The Effective Equations for Bianchi IX Loop Quantum Cosmology The Effective Equations for Bianchi IX Loop Quantum Cosmology Asieh Karami 1,2 with Alejandro Corichi 2,3 1 IFM-UMICH, 2 CCM-UNAM, 3 IGC-PSU Mexi Lazos 2012 Asieh Karami Bianchi IX LQC Mexi Lazos 2012

More information

CHAPTER 1 PRELIMINARIES

CHAPTER 1 PRELIMINARIES CHAPTER 1 PRELIMINARIES 1.1 Introduction The aim of this chapter is to give basic concepts, preliminary notions and some results which we shall use in the subsequent chapters of the thesis. 1.2 Differentiable

More information

BLACK HOLES IN LOOP QUANTUM GRAVITY. Alejandro Corichi

BLACK HOLES IN LOOP QUANTUM GRAVITY. Alejandro Corichi BLACK HOLES IN LOOP QUANTUM GRAVITY Alejandro Corichi UNAM-Morelia, Mexico ICGC 07, IUCAA, December 20th 2007 1 BLACK HOLES AND QUANTUM GRAVITY? Black Holes are, as Chandrasekhar used to say:... the most

More information

Quantum Symmetry Reduction for Diffeomorphism Invariant Theories of Connections

Quantum Symmetry Reduction for Diffeomorphism Invariant Theories of Connections PITHA 99/23 hep-th/9907042 arxiv:hep-th/9907042v1 7 Jul 1999 Quantum Symmetry Reduction for Diffeomorphism Invariant Theories of Connections M. Bojowald 1 and H.A. Kastrup 2 Institute for Theoretical Physics

More information

Self-adjoint extensions of symmetric operators

Self-adjoint extensions of symmetric operators Self-adjoint extensions of symmetric operators Simon Wozny Proseminar on Linear Algebra WS216/217 Universität Konstanz Abstract In this handout we will first look at some basics about unbounded operators.

More information